Proton coordination chemistry in pyrene-based anodes for ultralong-life aqueous proton batteries

Wenhui Yang a, Chensi Zhan a, Qiang Zhu b, Lei Liu a, Baiming Su a, Haoxiang Yu a, Liyuan Zhang a, Lei Yan *a and Jie Shu *a
aSchool of Materials Science and Chemical Engineering, Ningbo University, Ningbo, Zhejiang 315211, China. E-mail: yanlei@nbu.edu.cn; shujie@nbu.edu.cn
bMolecular Horizons and School of Chemistry and Molecular Bioscience, University of Wollongong, NSW 2522, Australia

Received 24th January 2025 , Accepted 23rd March 2025

First published on 25th March 2025


Abstract

Sustainable and safe aqueous proton batteries (APBs) have attracted significant attention owing to their unique “Grotthuss mechanism”. Although organic small molecules with stable and adjustable frameworks are promising electrode materials, their easy dissolution in electrolytes and unsatisfactory intrinsic conductivity hinder their broad application in APB devices. Herein, 2,7-diammonio-4,5,9,10-tetraone (PTO-NH3+) with stable intermolecular hydrogen-bond networks was designed via an in situ electrochemical reduction strategy. The optimized molecule structure endows low charge transport barriers, high chemical reactivity, and prominent charge affinity. The fast kinetics of proton coordination/de-coordination behavior in PTO-NH3+ electrodes is corroborated by ex situ characterization techniques and theoretical calculations. As a result, the robust four-step 4e H+ coordination with PTO-NH3+ electrode achieves an excellent rate performance (214.3 mA h g−1 at 0.05 A g−1 and112.9 mA h g−1 at 40 A g−1), along with a long lifespan (10[thin space (1/6-em)]000 cycles). These findings shed light on further avenues towards advanced proton batteries.


1. Introduction

With the contradiction between the unparalleled lack of fossil fuels and the increasing global energy demand intensifying, electrochemical energy storage devices, which provide a feasible approach to realizing efficient utilization of renewable energy sources such as solar and wind, have drawn widespread attention.1–4 Impressively, aqueous batteries stand out for abundant resources, intrinsic safety, and environmental friendliness due to their innoxious and apyrous electrolyte.5–9 Earth-abundant metal ions (such as Na+, K+, Mg2+, Ca2+, Zn2+, and Al3+)10–28 are selected as charge carriers for aqueous batteries. Nevertheless, owing to their larger ionic radius and charge number than Li+,29 it is challenging for these aqueous batteries to achieve competitive performance with lithium-ion batteries.

In view of the smallest ionic radius and eminent abundance in nature, proton (H+) with ultrafast diffusion kinetics and negligible cost30 is a considerably attractive substitute for metal cations in the construction of aqueous rechargeable batteries. Since the prototype of a proton battery was first reported by Sjödin's group in 2017,31 an increased number of proton storage materials have emerged, including metal oxides (e.g. MoO3, WO3, TiO2 and V2O5),32–35 Prussian blue analogues (PBAs) (e.g. Cu[Fe(CN)6]0.63·3.4H2O, Ni[Fe(CN)6]2/3·4H2O),36,37 organic solids (e.g. 3,4,9,10-perylenetetracarboxylic dianhydride [PTCDA],38 pyrene-4,5,9,10-tetraone [PTO],39 diquinoxalino [2,3-a:2′,3′-c] phenazine [HATN],40 and poly(3,4-ethylenedioxythiophene) [PEDOT]),41 and MXenes.42 Inorganic electrodes often suffer from poor cycle stability, low specific capacity and low redox potential. In this regard, organic electrode materials demonstrate significant advantages, such as structural versatility and functional tunability. However, organic-based proton batteries often suffer from fast capacity attenuation during repeated charge–discharge cycling because of the solubility of the active materials. Furthermore, currently reported organic electrodes for proton storage exhibit slow intermolecular ion diffusion kinetics, resulting in poor rate performance and low redox-active site utilization. Recently, some reports have suggested that intermolecular H-bonds can improve the solvation energy of organic materials, thus inhibiting their solubility.43–46 In addition, the weak intermolecular hydrogen bond interaction can form a horizontal two-dimensional extended superposition network, thereby improving cycle tolerance and ion transport performance of organic molecules.47–50 However, an in-depth investigation into the mechanism of hydrogen bond formation between organic electrodes during proton storage is still lacking.

Herein, a planar and highly-delocalized organic electrode material (PTO-NH3+) with stable intermolecular hydrogen-bond networks was synthesized by nitration and electrochemical reduction for aqueous proton storage. Benefiting from its overall molecular rigidity and facilitated aromaticity, the PTO-NH3+ exhibits rapid charge transport and an optimized energy bandgap, resulting in enhanced intrinsic conductivity, high redox activity, and superior electron affinity for proton storage. When served as an electrode material, it demonstrates fast coordination kinetics with an exceptional capacity as high as 214.3 mA h g−1 at 0.05 A g−1, splendid rate performance (40 A g−1), and long-term durability (≥10[thin space (1/6-em)]000 cycles) in 9.5 m H3PO4 electrolyte. Multiple ex situ characterization techniques further revealed that the highly reversible redox centers of PTO-NH3+ upon proton uptake/removal are C[double bond, length as m-dash]O groups, while the corresponding protonation pathway and structural evolution are validated by theoretical calculations. These breakthroughs pave the way for the design of outstanding organic electrode materials for APBs.

2. Results and discussion

The DNPT sample was straightforwardly synthesized via a one-step nitration reaction of pyrene-4,5,9,10-tetraone (PTO) (Fig. 1a). Fourier transform infrared (FT-IR) spectra of DNPT and PTO were collected, as illustrated in Fig. 1b. Examination of Fig. 1b shows that the peak at 1680 cm−1 is assigned to the C[double bond, length as m-dash]O bond. The asymmetric and symmetric stretching vibration peaks at 1531 and 1348 cm−1 can be attributed to the N–O bonds of the nitro group (–NO2). The Raman spectra of DNPT and PTO were also collected. As depicted in Fig. 1c, the –NO2 and C[double bond, length as m-dash]O peaks appear at 1349 and 1696 cm−1, respectively.51 The chemical structure of DNPT was verified by 1H and 13C nuclear magnetic resonance (NMR) (Fig. 1d and e). As shown in Fig. 1d, due to the high symmetry of the molecular structure, the peak at 8.91 ppm corresponds to the hydrogen atom (1) on the pyrene unit. Fig. 1e shows the 13C NMR diagram of DNPT, where the five peaks at 149.20, 126.48, 134.56, 136.35, and 174.73 ppm are associated with carbon atoms at 4, 1, 2, 3, and 5 sites, respectively. The morphology of the DNPT sample was observed by scanning electron microscopy (SEM) at low and high magnification (Fig. S1), revealing its stacked flake structure with a side length of 7–10 μm, and the transmission electron microscopy (TEM) image (Fig. S2) is consistent with this finding. X-ray diffraction (XRD) patterns were collected in Fig. S3, likewise corroborating the as-prepared sample is DNPT, while the sharp diffraction peaks demonstrate its fine crystallinity, which can be ascribed to the π–π stacking between pyrene units.52 The thermogravimetric (TG) curve (Fig. S4) was further employed to certify its thermodynamical stability.
image file: d5qi00269a-f1.tif
Fig. 1 Structural characterization of intermediates. (a) Synthetic process of DNPT. (b) FTIR spectra of PTO and DNPT. (c) Raman spectra of PTO and DNPT. (d) 1H NMR spectrum and (e) 13C NMR spectrum of DNPT (DMSO-d6).

The electrochemical properties of the DNPT electrode were estimated in a typical three-electrode system utilizing activated carbon as a counter electrode. 9.5 m H3PO4 was selected as an electrolyte owing to its prominent conductivity for protons, wherein protons have incomplete solvation shells that enable protons to directly interact with the electrode surface.53 Note that all potentials are relative to the Ag/AgCl reference electrode. The galvanostatic charge/discharge measurements in Fig. 2a show that DNPT can deliver a high initial discharge capacity of 1231.9 mA h g−1 and a charge capacity of 429.3 mA h g−1 at a current density of 0.05 A g−1, giving a coulombic efficiency (CE) of 34.8%. Such an initial irreversible capacity loss can be attributed to the conversion from the –NO2 group in DNPT to the –NH2 group.54 To gain insights into this reduction mechanism, the DNPT electrodes before/after the electrochemical reaction were further investigated by X-ray photoelectron spectrometry (XPS) to confirm the formation of PTO-NH2. The high-resolution N 1s XPS spectra are presented in Fig. 2b, indicating that the pristine DNPT electrode presented a characteristic peak of –NO2 at 406.1 eV, with a satellite peak at 400.5 eV. After discharge, a signal at 399.9 eV is attributed to the amino group (–NH2) emerges, while the –NO2 peak completely disappears. During the subsequent cycle, no trace of –NO2 is detected again, indicating an irreversible reduction from –NO2 to –NH2. This point was further confirmed by the FTIR spectra (Fig. 2c) and Raman spectra (Fig. S5) at different charge/discharge states. As shown in Fig. 2c, two characteristic –NO2 stretching bands at 1528 and 1346 cm−1 can be observed in the pristine DNPT electrode. After fully discharging to 0 V, the –NO2 peaks almost completely disappear. Meanwhile, new peaks appear at 3500–3300 cm−1. Among them, the peak at 3430 cm−1 belongs to the symmetric stretching vibration of free –NH2 (νN–H). The peaks at 3360 and 3235 cm−1 correspond to the associated –NH2 (νN–H⋯O and νN–H⋯N). The peak at 1586 cm−1 was assigned to the bending vibration of –NH2 (δN–H). The presence of these peaks indicates the formation of –NH2 groups during discharge. At the same time, the C[double bond, length as m-dash]O bonds form a hydrogen-bond coordination structure, which transforms into C[double bond, length as m-dash]O⋯H. When the DNPT electrode was fully charged to 1.0 V, the characteristic peaks of –NH2 remain, and no –NO2 signal is detected. Moreover, the difference can be also identified in the crystal structure of the DNPT electrode before/after discharge, which is reflected by the XRD pattern (Fig. 2d). After full discharge, the diffraction peaks of the DNPT electrode completely disappeared, and a new diffraction peak at 27.3° appeared, revealing the formation of a new crystal structure. After a full charge, an obvious peak at 28.6° can be observed, which corresponds to the π–π stacking structure of PTO-NH2. The morphological evolution of the DNPT electrode during the initial charging/discharging process was also reflected in the SEM images (Fig. S6). Based on the above analysis results, the DNPT was converted to PTO-NH2 through the in situ electrochemical reduction method. In acid electrolytes, PTO-NH2 mainly exists in the form of PTO-NH3+ due to hydrogen bond interactions, as illustrated in Fig. 2e. Furthermore, the structure of the intermolecular hydrogen-bond networks based on the –NH2 groups in the PTO-NH2 is shown in Fig. S7.


image file: d5qi00269a-f2.tif
Fig. 2 Electrochemical reduction mechanism of the DNPT electrode. (a) Voltage–capacity profile of the DNPT with marked points for the XPS tests and (b) corresponding ex situ N 1s XPS spectra. (c) FTIR spectra of the DNPT electrodes at different charge/discharge states. (d) XRD patterns of the DNPT electrodes at different charge/discharge states. (e) Synthetic process of PTO-NH3+.

The proton-storage metrics of the PTO-NH3+ electrode were evaluated by cyclic voltammetry (CV) and galvanostatic charge/discharge measurements. The CV curves of the PTO-NH3+ electrode at various scan rates from 0.2 to 1 mV s−1 are collected in Fig. 3a, which displays four pairs of symmetric redox peaks at 0.10/0.12 V, 0.24/0.26 V, 0.38/0.40 V and 0.62/0.64 V, indicating a four-step protonation/deprotonation electrochemistry. The corresponding peak shapes were virtually identical as the scan rates increased, revealing that the PTO-NH3+ electrode exhibits extraordinary electrochemical reversibility. Further investigation of proton storage kinetics is obtained according to the power-law relationship between the peak current (i) and scan rate (v): i = avb. As illustrated in Fig. 3b, the calculated b-values for the R1/O1, R2/O2, R3/O3, and R4/O4 redox peaks are 0.87/0.96, 0.90/1.02, 0.96/0.97, and 0.99/0.93, respectively, suggesting that the H+ uptake/removal process is dominated by capacitive behavior. The hydrogen-bond network between molecules plays a critical role in the energy storage process of organic batteries. The specific mechanisms are as follows: (1) organic small molecules form a network through hydrogen bonds, providing channels for proton conduction and reducing protonation activation energy. The dynamic nature of hydrogen bonds allows protons to transfer rapidly through the Grotthuss mechanism, thereby accelerating the electrochemical reactions on the electrode surface; (2) hydrogen-bond network enables proton conduction to occur primarily over short distances, reducing the need for long-range diffusion and creating localized regions of high proton concentration on the electrode surface, further minimizing the impact of diffusion limitations. Additionally, the corresponding contribution ratios at each scan rate were quantified using the Trasatti equation: i = k1v + k2v1/2. As exhibited in Fig. 3c and Fig. S8, the capacitive contribution increased gradually as the scan rate increased from 0.2 to 1 mV s−1. When the scan rate reached 1 mV s−1, a capacitive contribution as high as 97.2% was achieved, which was conducive to achieving excellent rate performance. In the same way, we depict the CV curves at different scan rates from 1 to 15 mV s−1 as well as relative b-values and contribution ratios (Fig. S9 and S10), which reflect conclusions in accordance with those mentioned above. The enhanced capacitive characteristics enabled the PTO-NH3+ electrode to exhibit rapid reaction kinetics.


image file: d5qi00269a-f3.tif
Fig. 3 Kinetics properties and electrochemical performance of the PTO-NH3+ electrode. (a) CV curves collected at various scan rates from 0.2 to 1 mV s−1. (b) Calculated b values. (c) Contribution ratio of the capacitive and diffusion-controlled charge versus scan rate. (d) Discharge/charge GITT curves for the PTO-NH3+ electrode at 0.1 A g−1. (e) H+ diffusion coefficient calculated from GITT. (f) Ex situ EIS plots at different states of charge. (g) GCD profiles at different current densities. (h) Rate capability of the PTO-NH3+ electrode. (i) Lifespan comparison of PTO-NH3+ with recently reported organic materials. (j) Cycle performance under a current density of 10 A g−1 for 10[thin space (1/6-em)]000 cycles.

To gain further insights into the intrinsic kinetics properties with respect to PTO-NH3+, the proton diffusion coefficient (DH+) and charge transfer resistance (Rct) were estimated by galvanostatic intermittent titration technique (GITT) and electrochemical impedance spectroscopy (EIS) measurements, respectively. The GITT curve of the PTO-NH3+ electrode is shown in Fig. 3d, and the DH+ results calculated according to the GITT curve are shown in Fig. 3e. As shown in Fig. 3e, the value of DH+ ranges from 10−10 to 10−6 cm2 s−1 over the whole discharge/charge process, confirming the superlative high-rate performance of PTO-NH3+ electrode. Interestingly, during the discharge process from 0.7 to 0.1 V, DH+ essentially maintains around 10−7 cm2 s−1, while subsequent deeper discharge to 0 V gives rise to the rapid decline of DH+. The same occurs during charging, where DH+ rapidly decreases as the potential reaches 0.8 V. To an extent, this phenomenon presumably stems from the critical concentration polarization at the end of the discharge/charge process. Targeting the discharge process, protons can readily coordinate with PTO-NH3+ in the beginning, in stark contrast to the almost filled reaction sites of PTO-NH3+ and thus generate strong electrostatic repulsion at the end of the discharge process, both of which render the diffusion of protons within the electrode more difficult.55,56 Besides, Fig. 3f compares the ex situ EIS plots at different states of charge, in which the diameter of the semicircle at the high-frequency region denotes Rct at the electrode/electrolyte interface. The exceedingly minor Rct values ranging from 2.58 to 2.79 Ω during the discharge/charge process indicate that the PTO-NH3+ electrode possesses low inherent resistance, which is attributed to the extensive π-electron delocalization via constructing hydrogen bond networks.52

The galvanostatic charge/discharge tests were then conducted at various current densities ranging from 0.05 A g−1 to 40 A g−1. As shown in Fig. 3g, the PTO-NH3+ electrode delivers considerable specific capacities of 214.3, 164.8, 152.2, 144.5, 140.8, 137.4, 134.2, 130.3, and 125.4 mA h g−1 at current densities of 0.05, 0.1, 0.2, 0.5, 1, 2, 5, 10, and 20 A g−1, respectively. Significantly, even when the current density increased 800-fold to 40 A g−1 (i.e., the battery was fully charged in only 10 s), the battery still maintained a decent capacity of 112.9 mA h g−1 (approximately 52.7% of the initial value), indicating its potential use in specialized fast-charging applications. In addition, as shown in Fig. 3h, at a low current density of 0.05 A g−1, the rapid decay of capacity in the initial 5 cycles, may be caused by the side reaction of remanent –NO2. However, when the current density was increased from 0.1 A g−1 to 40 A g−1, only 31% capacity attenuation occurred, indicating the ultrafast kinetics and superior rate performance of the PTO-NH3+ electrode. By contrast, this rate performance significantly overwhelms most organic materials reported for proton storage (Fig. 3i and Table S1).41,57–62 In addition to excellent rate capability, the PTO-NH3+ electrode also manifests promising cycling stability, as evidenced by its capacity retention rate of 85% after 450 cycles at 0.1 A g−1 (Fig. S12), and still maintains a capacity retention rate of 78% after 5000cycles at 5 A g−1, (Fig. S13). Impressively, even at an exceedingly high current density of 10 A g−1, the electrode retains 74.1% capacity over 10[thin space (1/6-em)]000 cycles with a CE of ≈100%, suggesting outstanding structural stability and promising prospects for practical applications of PTO-NH3+ electrodes.

The underlying causes of capacity fade were also thoroughly investigated. As shown in Fig. S14, the morphology of the electrode changed dramatically, with several floral clusters appearing after cycling, indicating the destabilization of the electrode surface environment. Furthermore, the disappearance of the unique diffraction peak at 28.6°, which corresponds to the π–π stacking structure of PTO-NH2 in Fig. S15, shows that the structural stability of the hydrogen-bond network is disrupted, fundamentally contributing to the decay of capacity.

Based on the high capacity and remarkable cycling performance at various current densities, it is evident that PTO-NH3+ exhibited exceptional electrochemical properties as an electrode material for H+ storage. Thus, the intermolecular hydrogen bond network makes the PTO-NH3+ electrode have higher structural stability and faster interfacial charge transfer rate, resulting in the PTO-NH3+ electrode having excellent cycle tolerance and rate performance. Meanwhile, Table S1 shows that the cycling stability of PTO-NH3+ is superior to that of most of the reported materials, indicating that hydrogen-bond networks offer significant advantages in suppressing electrode dissolution. Compared with COFs and polymer electrodes, hydrogen-bond networks provide a better balance between dissolution inhibition and electrochemical performance, particularly in high-rate, low-temperature, and flexible applications. These properties make hydrogen-bond network-based organic electrode materials highly promising for next-generation proton batteries.

To clarify the storage mechanism of H+, ex situ FTIR and XPS tests were applied to track the structural evolution of PTO-NH3+ electrode at selected discharge and charge states, as marked in Fig. 4a. In the FTIR spectra (Fig. 4b and Fig. S16), the characteristic signal of C[double bond, length as m-dash]O bonds in PTO-NH3+ at 1672 cm−1 gradually weakens upon the deepening of discharge. Meanwhile, new absorption peaks corresponding to C–O bonds and O–H stretch appear at 1415 cm−1 and around 3329 cm−1, respectively.43,56 Moreover, the intensity of these peaks (i.e., C–O and O–H absorption peaks) gradually increases during discharge. This phenomenon indicates that the C[double bond, length as m-dash]O bonds of PTO-NH3+ are the active sites and are involved in the storage of H+ ions. The reversible evolution of the O–H absorption peak also reflects the interaction of the hydrogen-bond networks with H+. Besides, during the subsequent recharging process, the intensity evolutions of the C[double bond, length as m-dash]O, C–O and O–H peaks exhibit the opposite trend to that of the discharge, suggesting that C–O bonds are reversibly transformed into C[double bond, length as m-dash]O bonds. Furthermore, ex situ XPS spectra of O 1s and C 1s (Fig. 4c and d) were obtained to confirm the coordination chemistry between the H+ ions and C[double bond, length as m-dash]O groups. The O 1s spectra show three peaks at 531.9, 533.1, and 534.1 eV at state I, attributed to C[double bond, length as m-dash]O, C–O, and O–H species, respectively. The existence of O–H species may be related to the hydrogen-bond networks between PTO-NH3+ molecules. During the discharge procedure (from state I to state VII), the intensity contribution of C[double bond, length as m-dash]O gradually declines, whereas that of C–O increases. The same trend occurs in the C 1s spectra, where a curve-fitted peak (288.1 eV) corresponding to the C[double bond, length as m-dash]O bond follows a decreasing path upon proton uptake. Also of note, both O 1s and C 1s spectra can revert to the initial states after recharging (from state VII to state XIII), which is consistent with the ex situ FTIR results. To summarize, carbonyl moieties (C[double bond, length as m-dash]O ↔ C–O/O–H) of the PTO-NH3+ electrode are uncovered as redox-active centers that prompt the reversible electrochemical reaction upon H+ (de)coordination.


image file: d5qi00269a-f4.tif
Fig. 4 Proton storage mechanism of PTO-NH3+ electrode. (a) GCD profile under a current density of 1 A g−1 with marked points for ex situ tests. (b) Overview of ex situ FTIR spectra. Ex situ XPS spectra of O 1s and C 1s during (c) discharging process and (d) charging process.

To further unravel the effect of the reduction from –NO2 to –NH2 on the PTO unit's electrochemistry and the proton uptake pathway in the PTO-NH3+ electrode, we conducted density functional theory (DFT) calculations with frontier molecular orbital theory and minimum energy principle.63–65 As shown in Fig. 5a, replacing –NO2 with –NH2 increases the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) from −8.089 and −4.771 eV to −6.036 and −3.545 eV, respectively. The corresponding energy gap (ΔE) of PTO-NH2 is reduced to 2.490 eV compared with 3.318 eV in DNPT. The narrower energy gap of PTO-NH2 is strongly correlated with the higher intrinsic electronic conductivity among organic materials (Table S2),66 which is beneficial to electron transfer when applied as electrode materials. The Gibbs free energy changes (ΔG) of PTO-NH3+ molecules incorporating various numbers of hydrogen atoms and their corresponding electrostatic potential distributions (ESPs) are given in Fig. 5b, from which we can see the electrostatic potential around functional C[double bond, length as m-dash]O was negative (nucleophilic center) while positive (electrophilic center) around –NH3+, indicating that H+ ions favor the active site of C[double bond, length as m-dash]O. During the protonation process, the protons successively coordinate with C[double bond, length as m-dash]O bonds in a four-step reaction procedure to form PTO-NH3+-H (−0.6433 eV), PTO-NH3+-2H (−0.7524 eV), PTO-NH3+-3H (−1.9777 eV), and PTO-NH3+-4H (−2.6276 eV), which conforms to the four pairs of redox peaks observed in CV curves. When the PTO-NH3+-4H molecular state is achieved, the continuous blue region on the surface of C[double bond, length as m-dash]O also almost becomes the red region, indicating that the discharge process is complete. Furthermore, the Gibbs free energy changes for incorporating hydrogen atoms are negative, demonstrating that the reaction should occur spontaneously. To gain more insight into the nature of the bonds formed between C[double bond, length as m-dash]O and hydrogen atoms, charge-density difference distributions were also estimated. As illustrated in Fig. 5c, the electron density change occurred mainly around the active site of C[double bond, length as m-dash]O, which agrees with the conclusions drawn from the ESP distribution. According to the Grotthuss mechanism, the transfer of free protons occurs directly in the adjacent hydrogen-bond networks of PTO-NH3+ molecules.


image file: d5qi00269a-f5.tif
Fig. 5 DFT calculations. (a) HOMO–LUMO plots of DNPT and PTO-NH2, together shown are their corresponding values and gaps in the unit of eV. (b) Charge density difference for PTO-NH3+, which possesses (i–iv) 1, 2, 3, and 4 hydrogen atoms, respectively. (c) Gibbs free energy change (ΔG) of PTO-NH3+ molecules reacted with different hydrogen atoms, together shown are their corresponding ESP images with a color scale ranging from −50 (blue) to 50 (red) a.u.

3. Conclusions

In summary, a planar organic electrode PTO-NH3+, which forms intermolecular hydrogen bonds through electron delocalization polarization between C[double bond, length as m-dash]O and NH2 has been demonstrated for splendid proton storage. Furthermore, the electrochemical mechanism of the PTO-NH3+ electrode was verified to be H+ coordination/de-coordination chemistry using ex situ characterization techniques and DFT calculations. The C[double bond, length as m-dash]O groups with negative ESP are the active sites for H+ uptake, and a four-step reversible structural evolution occurs from PTO-NH3+ to PTO-NH3+-H, PTO-NH3+-2H, PTO-NH3+-3H, and PTO-NH3+-4H, respectively. In addition, the introduction of –NH3+ narrows the energy gap, which facilitates electronic conductivity and charge affinity. Owing to the fast kinetics of H+ coordination/de-coordination behavior, the PTO-NH3+ electrode not only achieves an outstanding proton-storage capacity of 214.3 mA h g−1 at 0.05 A g−1, but also represents excellent rate performance of 112.9 mA h g−1 at 40 A g−1 and long-term cycle durability (capacity retention of 74.1% over 10[thin space (1/6-em)]000 cycles at 10 A g−1) in the 9.5 m H3PO4 electrolyte. This work provides a novel strategy for the engineering of organic small molecules for ultrafast proton storage and lays a solid foundation for further advancement of aqueous proton electrochemistry.

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of interest

The authors declare that they have no known competing financial interests or personal relationships that could influence the work reported in this paper.

Acknowledgements

This work was financially supported by the Natural Science Foundation of Zhejiang Province (LY23B030005), National Natural Science Foundation of China (22209082), and Natural Science Foundation of Ningbo Municipality (2023J098).

References

  1. M. R. Lukatskaya, B. Dunn and Y. Gogotsi, Multidimensional materials and device architectures for future hybrid energy storage, Nat. Commun., 2016, 7, 12647 Search PubMed.
  2. P. Simon and Y. Gogotsi, Perspectives for electrochemical capacitors and related devices, Nat. Mater., 2020, 19, 1151–1163 Search PubMed.
  3. D. Chao, W. Zhou, F. Xie, C. Ye, H. Li, M. Jaroniec and S. Qiao, Roadmap for advanced aqueous batteries: From design of materials to applications, Sci. Adv., 2020, 6, eaba4098 Search PubMed.
  4. M. Shi, P. Das, Z. Wu, T. Liu and X. Zhang, Aqueous organic batteries using the proton as a charge carrier, Adv. Mater., 2023, 35, 2302199 Search PubMed.
  5. S. Chen, M. Zhang, P. Zou, B. Sun and S. Tao, Historical development and novel concepts on electrolytes for aqueous rechargeable batteries, Energy Environ. Sci., 2022, 15, 1805–1839 Search PubMed.
  6. Z. Ju, Q. Zhao, D. Chao, Y. Hou, H. Pan, W. Sun, Z. Yuan, H. Li, T. Ma, D. Su and B. Jia, Energetic aqueous batteries, Adv. Energy Mater., 2022, 12, 2201074 Search PubMed.
  7. Z. Tie, L. Liu, S. Deng, D. Zhao and Z. Niu, Proton insertion chemistry of a zinc-organic battery, Angew. Chem., Int. Ed., 2020, 59, 4920–4924 Search PubMed.
  8. H. Zhang, X. Liu, H. Li, I. Hasa and S. Passerini, Challenges and strategies for high-energy aqueous electrolyte rechargeable batteries, Angew. Chem., Int. Ed., 2021, 60, 598–616 Search PubMed.
  9. Y. Shang, N. Chen, Y. Li, S. Chen, J. Lai, Y. Huang, W. Qu, F. Wu and R. Chen, An “ether-in-water” electrolyte boosts stable interfacial chemistry for aqueous lithium-ion batteries, Adv. Mater., 2020, 32, 2004017 Search PubMed.
  10. Z. Tie, S. Deng, H. Cao, M. Yao, Z. Niu and J. Chen, A symmetric all-organic proton battery in mild electrolyte, Angew. Chem., Int. Ed., 2022, 61, e202115180 Search PubMed.
  11. H. Cui, D. Zhang, Z. Wu, J. Zhu, P. Li, C. Li, Y. Hou, R. Zhang, X. Wang, X. Jin, S. Bai and C. Zhi, Tailoring hydroxyl groups of organic phenazine anodes for high-performance and stable alkaline batteries, Energy Environ. Sci., 2024, 17, 114–122 Search PubMed.
  12. Y. Lin, H. Cui, C. Liu, R. Li, S. Wang, G. Qu, Z. Wei, Y. Yang, Y. Wang, Z. Tang, H. Li, H. Zhang, C. Zhi and H. Lv, A covalent organic framework as a long-life and high-rate anode suitable for both aqueous acidic and alkaline batteries, Angew. Chem., Int. Ed., 2023, 62, e202218745 Search PubMed.
  13. F. Ye, Q. Liu, H. Dong, K. Guan, Z. Chen, N. Ju and L. Hu, Organic zinc-ion battery: planar, π-conjugated quinone-based polymer endows ultrafast ion diffusion kinetics, Angew. Chem., Int. Ed., 2022, 61, e202214244 Search PubMed.
  14. X. Wang, T. Li, X. Zhang, Y. Wang, H. Li, H. Li, G. Zhao and C. Han, High-performance magnesium/sodium hybrid ion battery based on sodium vanadate oxide for reversible storage of Na+ and Mg2+, J. Energy Chem., 2024, 96, 79–88 Search PubMed.
  15. X. Ma, B. Zhao, H. Liu, J. Tan, H. Li, X. Zhang, J. Diao, J. Yue, G. Huang, J. Wang and F. Pan, H2O-Mg2+ waltz-like shuttle enables high-capacity and ultralong-life magnesium-ion batteries, Adv. Sci., 2024, 11, 2401005 Search PubMed.
  16. R. P. Jadav, D. Singh, R. Ahuja and Y. Sonvane, Fluorine-terminated Mxene as an anode material for dual-ion (Ca2+/Mg2+) batteries with rapid diffusion mobility, J. Phys. Chem. C, 2024, 128, 13539–13549 Search PubMed.
  17. R. L. Streng, S. Reiser, S. Wager, N. Pommer and A. S. Bandarenka, A fast and highly stable aqueous calcium-ion battery for sustainable energy storage, ChemSusChem, 2024, 18, e202401469 Search PubMed.
  18. L. Yan, Q. Zhu, Y. Qi, J. Xu, Y. Peng, J. Shu, J. Ma and Y. Wang, Towards high-performance aqueous zinc batteries via a semi-conductive bipolar-type polymer cathode, Angew. Chem., Int. Ed., 2022, 61, e202211107 Search PubMed.
  19. G. Yoo, Y. Jo and G. An, Multifunctional zinc vanadium oxide layer on metal anodes via ultrathin surface coating for enhanced stability in aqueous zinc-ion batteries, ACS Energy Lett., 2024, 9, 5955–5965 Search PubMed.
  20. V. P. H. Radhakantha, S. Pradhan and A. J. Bhattacharyya, Exploring aluminum-ion (Al3+) insertion in ammonium vanadium bronze (NH4V4O10) for aqueous aluminum-ion rechargeable batteries, J. Phys. Chem. C, 2024, 128, 20025–20034 Search PubMed.
  21. Y. Wang, T. Wu, Y. Lu, W. Zhang and Z. Li, In situ synthesis of MoO3 by surface oxidation of Mo2C (MXene) for stable near-surface reactions in aqueous aluminum-ion battery, Angew. Chem., Int. Ed., 2024, 64, e202416032 Search PubMed.
  22. Z. Dai, R. Chanajaree, C. Yang, X. Zhang, M. Okhawilai, P. Pattananuwat, X. Zhang, G. He and J. Qin, Dual-functional additives boost zinc-ion battery electrolyte over wide temperature range, Energy Mater. Adv., 2025, 6, 0139 Search PubMed.
  23. Z. Li, T. T. Beyene, K. Zhu and D. Cao, Realizing fast plating/stripping of high-performance Zn metal anode with a low Zn loading, J. Met., Mater. Miner., 2024, 34, 2009 Search PubMed.
  24. K. Lolupiman, J. Cao, D. Zhang, C. Yang, X. Zhang and J. Qin, A review on the development of metals-doped Vanadium oxides for zinc-ion battery, J. Met., Mater. Miner., 2024, 34, 2084 Search PubMed.
  25. P. Sintipditsakul, C. Yang, Z. Dai, N. Kiatwisarnkij, K. Lolupiman, P. Woottapanit, X. Zhang, P. Wangyao and J. Qin, Construction of artificial interface layer in the fly ash suspension for durable Zn anode, ACS Appl. Energy Mater., 2025, 8, 1766–1775 Search PubMed.
  26. M. Songpanit, K. Boonyarattanakalin, W. Pecharapa and W. Mekprasart, ZnO nanostructures synthesized by one-step sol-gel process using different zinc precursors, J. Met., Mater. Miner., 2024, 34, 1968 Search PubMed.
  27. C. Yang, P. Woottapanit, S. Geng, R. Chanajaree, K. Lolupiman, W. Limphirat, X. Zhang and J. Qin, Biomimetic inorganic-organic protective layer for highly stable and reversible zn anodes, ACS Energy Lett., 2024, 10, 337–344 Search PubMed.
  28. C. Yang, P. Woottapanit, S. Geng, R. Chanajaree, Y. Shen, K. Lolupiman, W. Limphirat, T. Pakornchote, T. Bovornratanaraks, X. Zhang, J. Qin and Y. Huang, A multifunctional quasi-solid-state polymer electrolyte with highly selective ion highways for practical zinc ion batteries, Nat. Commun., 2025, 16, 183 Search PubMed.
  29. T. Xu, D. Wang, Z. Li, Z. Chen, J. Zhang, T. Hu, X. Zhang and L. Shen, Electrochemical proton storage: from fundamental understanding to materials to devices, Nano-Micro Lett., 2022, 14, 126 Search PubMed.
  30. S. Wu, H. Guo and C. Zhao, Challenges and opportunities for proton batteries: from electrodes, electrolytes to full-cell applications, Adv. Funct. Mater., 2024, 34, 2405401 Search PubMed.
  31. R. Emanuelsson, M. Sterby, M. Stromme and M. Sjödin, An all-organic proton battery, J. Am. Chem. Soc., 2017, 139, 4828–4834 Search PubMed.
  32. Y. Wu, W. Liu, Z. Zhang, Y. Zheng, X. Fu, J. Lu, S. Cheng, J. Su and Y. Gao, Defect-rich MoO3 nanobelts for ultrafast and wide-temperature proton battery, Energy Storage Mater., 2023, 61, 102849 Search PubMed.
  33. K. Roy, P. K. Sarga, S. Bhattacharyya and S. K. Das, Unveiling the role of electrolyte to promote enhanced proton storage stability in WO3 for energy storage, Mater. Chem. Phys., 2024, 320, 129461 Search PubMed.
  34. C. Geng, T. Sun, Z. Wang, J. Wu, Y. Gu, H. Kobayashi, P. Yang, J. Hai and W. Wen, Surface-induced desolvation of hydronium ion enables anatase TiO2 as an efficient anode for proton batteries, Nano Lett., 2021, 21, 7021–7029 Search PubMed.
  35. Q. Zhao, L. Liu, J. Yin, J. Zheng, D. Zhang, J. Chen and L. A. Archer, Proton intercalation/de-intercalation dynamics in vanadium oxides for aqueous aluminum electrochemical cells, Angew. Chem., Int. Ed., 2020, 59, 3048–3052 Search PubMed.
  36. X. Wu, J. Hong, W. Shin, L. Ma, T. Liu, X. Bi, Y. Yuan, Y. Qi, T. W. Surta, W. Huang, J. Neuefeind, T. Wu, P. A. Greaney, J. Lu and X. Ji, Diffusion-free Grotthuss topochemistry for high-rate and long-life proton batteries, Nat. Energy, 2019, 4, 123–130 Search PubMed.
  37. X. Wu, S. Qiu, Y. Xu, L. Ma, X. Bi, Y. Yuan, T. Wu, R. Shahbazian-Yassar, J. Lu and X. Ji, Hydrous nickel-iron Turnbull's blue as a high-rate and low-temperature proton electrode, ACS Appl. Mater. Interfaces, 2020, 12, 9201–9208 Search PubMed.
  38. X. Wang, C. Bommier, Z. Jian, Z. Li, R. S. Chandrabose, I. A. Rodríguez-Pérez, P. A. Greaney and X. Ji, Hydronium-ion batteries with perylenetetracarboxylic dianhydride crystals as an electrode, Angew. Chem., Int. Ed., 2017, 56, 2909–2913 Search PubMed.
  39. Z. Guo, J. Huang, X. Dong, Y. Xia, L. Yan, Z. Wang and Y. Wang, An organic/inorganic electrode-based hydronium-ion battery, Nat. Commun., 2020, 11, 959 Search PubMed.
  40. Y. Ma, Y. Wei, W. Han, Y. Tong, A. Song, J. Zhang, H. Li, X. Li and J. Yang, Proton intercalation/de-intercalation chemistry in phenazine-based anode for hydronium-ion batteries, Angew. Chem., Int. Ed., 2023, 62, e202314259 Search PubMed.
  41. M. Zhu, L. Zhao, Q. Ran, Y. Zhang, R. Peng, G. Lu, X. Jia, D. Chao and C. Wang, Bioinspired catechol-grafting PEDOT cathode for an all-polymer aqueous proton battery with high voltage and outstanding rate capacity, Adv. Sci., 2022, 9, 2103896 Search PubMed.
  42. J. Zhu, X. Li, B. Hu, S. Ge and J. Xu, Low-temperature-tolerant aqueous proton battery with porous Ti3C2Tx MXene electrode and phosphoric acid electrolyte, Batteries, 2024, 10, 207 Search PubMed.
  43. Z. Song, L. Miao, Y. Lv, L. Gan and M. Liu, NH4+ charge carrier coordinated H-bonded organic small molecule for fast and superstable rechargeable zinc batteries, Angew. Chem., Int. Ed., 2023, 62, e202309446 Search PubMed.
  44. C. Ding, Y. Zhao, W. Yin, F. Kang, W. Huang and Q. Zhang, Regulating intermolecular hydrogen bonds in organic cathode materials to realize ultra-stable, flexible and low-temperature aqueous zinc-organic batteries, Angew. Chem., Int. Ed., 2024, 64, e202417988 Search PubMed.
  45. Z. Lin, H. Shi, L. Lin, X. Yang, W. Wu and X. Sun, A high capacity small molecule quinone cathode for rechargeable aqueous zinc-organic batteries, Nat. Commun., 2021, 12, 4424 Search PubMed.
  46. S. Zheng, D. Shi, T. Sun, L. Zhang, W. Zhang, Y. Li, Z. Guo, Z. Tao and J. Chen, Hydrogen bond networks stabilized high-capacity organic cathode for lithium-ion batteries, Angew. Chem., Int. Ed., 2023, 62, e202217710 Search PubMed.
  47. S. Wu, M. Taylor, H. Guo, S. Wang, C. Han, J. Vongsvivut, Q. Meyer, Q. Sun, J. Ho and C. Zhao, A high-capacity benzoquinone derivative anode for all-organic long-cycle aqueous proton batteries, Angew. Chem., Int. Ed., 2024, 63, e202412455 Search PubMed.
  48. Y. Wang, C. Wang, W. Wang, Y. Zhang, Z. Guo, J. Huang, L. Yan, J. Ma and Y. Wang, Organic hydronium-ion battery with ultralong life, ACS Energy Lett., 2023, 8, 1390–1396 Search PubMed.
  49. X. Yan, F. Wang, X. Su, J. Ren, M. Qi, P. Bao, W. Chen, C. Peng and L. Chen, A redox-active covalent organic framework with highly accessible aniline-fused quinonoid units affords efficient proton charge storage, Adv. Mater., 2023, 35, 2305037 Search PubMed.
  50. C. Wang, D. He, H. Wang, J. Guo, Z. Bao, Y. Feng, L. Hu, C. Zheng, M. Zhao, X. Wang and Y. Wang, Symmetrical design of biphenazine derivative anode for proton ion batteries with high voltage and long-term cycle stability, Adv. Sci., 2024, 11, 2401314 Search PubMed.
  51. X. Liu and Z. Ye, Nitroaromatics as high-energy organic cathode materials for rechargeable alkali-ion (Li+, Na+, and K+) batteries, Adv. Energy Mater., 2021, 11, 2003281 Search PubMed.
  52. R. Wang, J. He, C. Yan, R. Jing, Y. Zhao, J. Yang, M. Shi and X. Yan, A long-range planar polymer with efficient π-electron delocalization for superior proton storage, Adv. Mater., 2024, 36, 2402681 Search PubMed.
  53. H. Jiang, W. Shin, L. Ma, J. Hong, Z. Wei, Y. Liu, S. Zhang, X. Wu, Y. Xu, Q. Guo, M. A. Subramanian, W. F. Stickle, T. Wu, J. Lu and X. Ji, A high-rate aqueous proton battery delivering power below-78 °C via an unfrozen phosphoric acid, Adv. Energy Mater., 2020, 10, 2000968 Search PubMed.
  54. Z. Chen, H. Su, P. Sun, P. Bai, J. Yang, M. Li, Y. Deng, Y. Li, Y. Gen and Y. Xu, A nitroaromatic cathode with an ultrahigh energy density based on six-electron reaction per nitro group for lithium batteries, Proc. Natl. Acad. Sci. U. S. A., 2022, 119, e2116775119 Search PubMed.
  55. J. Zheng, W. Shi, M. Gu, J. Xiao, P. Zuo, C. Wang and J. Zhang, Electrochemical kinetics and performance of layered composite cathode material Li[Li0.2Ni0.2Mn0.6]O2, J. Electrochem. Soc., 2013, 160, A2212–A2219 Search PubMed.
  56. Z. Su, J. Tang, J. Chen, H. Guo, S. Wu, S. Yin, T. Zhao, C. Jia, Q. Meyer, A. Rawal, J. Ho, Y. Fang and C. Zhao, Co-insertion of water with protons into organic electrodes enables high-rate and high-capacity proton batteries, Small Struct., 2023, 4, 2200257 Search PubMed.
  57. J. Qiao, M. Qin, Y. Shen, J. Cao, Z. Chen and J. Xu, A rechargeable aqueous proton battery based on a dipyridophenazine anode and an indium hexacyanoferrate cathode, Chem. Commun., 2021, 57, 4307–4310 Search PubMed.
  58. Y. Dai, X. Yan, J. Zhang, C. Wu, Q. Guo, J. Luo, M. Hu and J. Yang, High-capacity proton battery based on π-conjugated N-containing organic compound, Electrochim. Acta, 2023, 442, 141870 Search PubMed.
  59. M. Shi, J. He, Y. Zhao, L. Zhao, K. Dai and C. Yan, In situ Raman investigation and application of phenazine-based organic electrode in aqueous proton batteries, Mater. Des., 2022, 222, 111043 Search PubMed.
  60. M. Shi, R. Wang, L. Li, N. Chen, P. Xiao, C. Yan and X. Yan, Redox-active polymer integrated with Mxene for ultra-stable and fast aqueous proton storage, Adv. Funct. Mater., 2023, 33, 2209777 Search PubMed.
  61. X. Wang, J. Zhou and W. Tang, Poly(dithieno 3,2-b:2′,3′-d pyrrole) twisting redox pendants enabling high current durability in all-organic proton battery, Energy Storage Mater., 2021, 36, 1–9 Search PubMed.
  62. G. Zhao, X. Yan, Y. Dai, J. Xiong, Q. Zhao, X. Wang, H. Yu, J. Gao, N. Zhang, M. Hu and J. Yang, Searching high-potential dihydroxynaphthalene cathode for rocking-chair all-organic aqueous proton batteries, Small, 2024, 20, 2306071 Search PubMed.
  63. Y. Cheng, X. Liu, Y. Guo, G. Dong, X. Hu, H. Zhang, X. Xiao, Q. Liu, L. Xu and L. Mai, Monodispersed sub-1 nm inorganic cluster chains in polymers for solid electrolytes with enhanced Li-ion transport, Adv. Mater., 2023, 35, 2303226 Search PubMed.
  64. L. Du, B. Zhang, W. Deng, Y. Cheng, L. Xu and L. Q. Mai, Hierarchically self-assembled MOF network enables continuous ion transport and high mechanical strength, Adv. Energy Mater., 2022, 12, 2200501 Search PubMed.
  65. Z. Jiang, S. Wang, X. Chen, W. Yang, X. Yao, X. Hu, Q. Han and H. Wang, Tape-casting Li0.34La0.56TiO3 ceramic electrolyte films permit high energy density of lithium-metal batteries, Adv. Mater., 2020, 32, 1906221 Search PubMed.
  66. J. He, M. Shi, H. Wang, H. Liu, J. Yang, C. Yan, J. Zhao, J.-L. Yang and X.-L. Wu, Ladder-type redox-active polymer achieves ultra-stable and fast proton storage in aqueous proton batteries, Angew. Chem., Int. Ed., 2024, 63, e202410568 Search PubMed.

Footnotes

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5qi00269a
These authors contributed equally to this work.

This journal is © the Partner Organisations 2025
Click here to see how this site uses Cookies. View our privacy policy here.